Sei sulla pagina 1di 61

Third Year Dynamics Lecture Notes

Michael Zaiser

Contents
1 Introduction

1.1

Prerequisites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.2

Course outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.3

Use of these notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.4

Basic concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2 The theory of systems with one degree of freedom

2.1

Lumping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2.2

Free vibration of a SDF system with no damping . . . . . . . . . . . . . . . . . .

2.2.1

Equation of motion approach . . . . . . . . . . . . . . . . . . . . . . . . .

2.2.2

Energy approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Free vibration of a SDF system with damping . . . . . . . . . . . . . . . . . . . .

10

2.3.1

Viscous damping and solution of the damped motion equation . . . . . . .

10

2.3.2

Canonical form of the damped free vibration equation . . . . . . . . . . .

14

2.3.3

Rotational vibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

16

Forced vibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

19

2.4.1

Periodic forcing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

19

2.4.2

Force transmission and vibration damping . . . . . . . . . . . . . . . . . .

22

2.3

2.4

2.4.3

Moving boundary condition and amplitude transmission . . . . . . . . . .

23

2.4.4

Rotating out of balance rotor . . . . . . . . . . . . . . . . . . . . . . . . .

25

2.4.5

Non-periodic forcing . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

27

2.4.6

Shock damping and shock response spectrum . . . . . . . . . . . . . . . .

30

3 The theory of systems with many degrees of freedom

33

3.1

Free vibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

33

3.2

Eigenvalue problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

37

3.3

Forced vibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

38

3.4

Orthogonal modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

39

4 Applications

42

4.1

Accelerometer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

42

4.2

Shaft whirling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

44

4.3

Beating . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

47

4.4

Anti-resonance and vibration absorbers . . . . . . . . . . . . . . . . . . . . . . .

52

4.5

Out of balance in reciprocating internal combustion engines . . . . . . . . . . . . .

53

ii

1 Introduction
1.1 Prerequisites
You have already studied some freely vibrating systems last year, and we will be building on this
knowledge. You will need to draw on the material you studied both in dynamics, mathematics and
solid mechanics. As the course starts, make sure you know how to draw free body diagrams (FBDs)
and extract the equations of motion from these diagrams. You have studied the theory of second
order ordinary differential equations with constant coefficients in second year maths. We will make
a lot of use of the results, so revise them! I will be going over some of the material in the first few
lectures, but it will only be to refresh your memory and to define my own notations, so dont rely on
learning it for the first time here. As a test, before you start the course you should be able to solve
x + 5x + 4x = exp(3t)

(1)

with initial conditions


x(0) = 0 ,

x(0)
=1 .

(2)

Remember you will need to find the two complementary functions and the particular integral and
then use the initial conditions to determine the constants in the general solution. Sound familiar?
Could you solve the same equation if the right hand side was sin(3t)? You will also need to know
how to use matrices to represent systems of equations and know the meaning and use of eigenvalues
and eigenvectors. To deal with the problems in tutorial and exam questions, you also should recall
what you have learned about rotational motion, moments of inertia and beam flexure.

1.2 Course outline


The focus of the course is dynamic vibration. You can find more details and a lecture by lecture
breakdown in the syllabus, available from the dynamics web page. We will study the theory of
vibration of single and multiple degree of freedom systems with and without forcing. With this
theoretical work, we will be able to look at a range of applications. These will include shaft whirling,
vibration dampers, balancing in internal combustion engines and vibration measuring devices.
We will look first at single degree of freedom (SDF) systems in some detail because they exhibit
many of the features of more complicated systems while remaining relatively easy to analyze mathematically. After reviewing free body diagrams and clarifying notation, we will consider a freely
vibrating SDF system with and without damping. The equation of motion for the system follows
from the free body diagram. It is a second order, linear, homogeneous ordinary differential equation
with constant coefficients and it can be tackled by methods which will already be familiar from previous courses. We will look at the physical significance of the solution and see how the parameters
of the system determine its vibrational characteristics. Having understood the behaviour of freely
1

vibrating systems, we will introduce a forcing term and consider its effects. Some new mathematical techniques will be introduced here. We will first look at systems forced by a periodic force
and then at the more difficult general case. At this point we will develop a general solution for all
SDF systems, free or forced and damped or undamped, and discuss the different properties of each
system and the methods used to analyse them.
The second part of the module will address multiple degree of freedom (MDF) systems. We will
see that the problem is similar to an SDF system but with some added complications. We find that
the mathematical analysis is similar in SDF and MDF systems if we replace the mass and stiffness
constants with matrices. We will look at how to obtain solutions of the matrix equations and what
these solutions tell us about the behaviour of the system. We will see that in many cases we can
make a coordinate transformation that will allow us to solve the MDF system as a series of SDF
systems. The equations for MDF systems are often very hard to solve. In practice they are usually
solved numerically, and we will look at some of the methods that are used to do this.
Along with the mathematical theory, we will look at some applications to engineering systems. We
will find in most instances that we must simplify the true situation but that the answers we get from
the theory agree quite well with experiments. We will look at shaft whirling, IC engine balancing, rotating out-of-balance systems, vibration absorbers, anti-resonance devices, accelerometers,
torsional and beam systems. Numerical methods are becoming the dominant method of analysing
vibrating systems and we will spend some time using MATLAB to explore how the methods work
and what we need to be able to do to use them. It is a common misconception that numerical methods are easy to use. We will see that it is indeed almost too easy to use numerical methods to get
answers, but rather more difficult to use them properly and get valid answers.
At the end of the course we will look at what we have achieved and hopefully conclude that we can
analyse many interesting systems that are important in engineering. We will also discuss what we
have not yet studied and see the limitations of our theories.

1.3 Use of these notes


I suggest that you will find the material easier to follow if you look at the relevant sections of these
notes before the lectures. Do the exercises in the notes as we go along; they will help your understanding. You need to be active in your study by trying out the techniques on problems yourself.
Passively reading the notes will not get you very far.
It should really go without saying that these notes are in no way a replacement for the lectures, and
not only because they are incomplete. The notes complement the lectures and relieve you of the
tedious, error-prone task of copying chunks of algebra from the board.

1.4 Basic concepts


We will be building on the work from previous courses. You should already be familiar with the
basic ideas of freely-vibrating systems. In particular, you ought to know the meaning of period and
frequency of vibration, inertial forces, spring stiffness, moments of inertia, beam bending, complementary functions and particular integrals. You should also be able to draw free body diagrams
correctly. We will revise free body diagrams before starting the course proper.
A free body diagram shows the real and inertial forces that act on each mass in a system. Once a
free body diagram has been constructed, the equation of motion can often be written down. When
constructing the free body diagram, it is important to bear in mind the direction of each force in the
system.

x
k
1

k
2

m
2

k
3

Figure 1: An illustrative two-degree of freedom system. The masses are free to move in the horizontal direction only. When x1 = 0 and x2 = 0 the system is in static equilibrium
The best way to see how this works is to consider an example. In Figure 1 the two objects with
masses m1 and m2 are free to move in the horizontal direction. However, they are not free to move
in the vertical direction, an example of a constraint. The masses can move independently, so two
coordinates are needed to specify their position. The system therefore has two degrees of freedom.
The three springs in Figure 1 have stiffness constants k1 , k2 and k3 and like all springs we will
consider they are linear. If a linear spring is extended by a distance x, it exerts a force kx where
k is its stiffness. From this information, we can draw the free body diagram for the system. As a
reference configuration we always chose the static equilibrium configuration of the system. Figure 2
shows the system after the masses have been moved by (arbitrary) distances of x1 and x2 out of this
configuration. If the distance between the masses in the equilibrium configuration is l, the distance
between them after they have moved is l x1 + x2 . The extension of the central spring is therefore
x2 x1 . The forces due to the springs can now be deduced. Figure 3 shows the forces acting on each
mass, including the inertial forces. Note that the final spring is compressed when x2 is positive and
so the force must be directed to the left. We choose to show this by drawing the arrow to the right
and including a minus sign. Equivalently, we could have kept the force as positive and directed the
arrow to the left.

l
x

l-x 1+ x

Figure 2: Distances when the masses have moved.

k 1x

k 2(x 2-x 1)
1

m 1 &x & 1
m

-k 3x
2

&x & 2

Figure 3: The free-body diagram for the illustrative system.


The direction of the inertial forces is important, and can be determined by considering the effect
of the inertia of the mass. If x1 is positive, as it is shown in Figure 2, then the mass m1 has been
moved to the right. If we accelerate a mass in the direction of x1 to the right, its inertia acts to
oppose this and so the inertial force is directed to the left. The same argument applies to the mass
m2 . Hence, the direction of the inertial forces must always be taken opposite to the direction of the
respective coordinate. The equations of motion for the two mass system can now be written down
using DAlemberts principle that the sum of the forces on each body must be zero in dynamic
equilibrium. We obtain one equation for each mass, and they are given by
m1 x1 + k1 x1 k2 (x2 x1 ) = 0 ,
m2 x2 + k2 (x2 x1 ) + k3 x2 = 0 .
Exercise: 1
It is a useful exercise at this point to redraw the system of Figure 1 with a different
coordinate system. Label the system with coordinates x3 and x4 so that x3 is just
the same as x1 but x4 runs in the opposite direction to x2 . Draw the free body
diagram and deduce the equations of motion. When you have done this, substitute
x3 = x1 and x4 = x2 and check you have the same equation of motion as derived
above. Tip: You may find you need to multiply one of the equations by -1 to make
it the same as equation (3).

(3)

What can we conclude from this brief exercise? Minus signs can be a headache when we come to
figure out the directions of forces, but actually the mathematics takes care of all this for us! If we
set up a coordinate system so that things work properly for positive displacements, everything will
also work for negative displacements. Therefore, when setting up a FBD, just specify a direction for
the displacement of a body (and stick to it!), and draw in the correctly directed spring force which
results from this displacement. If in reality the body moves the other way, it doesnt matter because
the maths will sort out the appropriate sign on the force.

2 The theory of systems with one degree of freedom


Systems with one degree of freedom are very important and will be the starting point for our study
of vibration. We will find that there are many interesting engineering problems that can be modelled
well by single degree of freedom (SDF) systems. Other more complicated systems will need extra
degrees of freedom, but we will discover that many of the ideas and techniques we develop for
SDF systems will be easily extendible to many degree of freedom (MDF) systems, although the
mathematical effort needed to solve them will be greater.

2.1 Lumping
Before we start to analyse a SDF system, let us consider for a moment what we are really thinking
about when we draw our initial diagram for a system. In reality, we can never cover all the complicated aspects of a real system. For instance, consider a mass supported by a spring in the absence
of gravity. This appears to be a very simple dynamic system to study, and we might draw a diagram
like that shown in Figure 4.
Remember there is no gravity so the equilibrium position of the mass is at the unextended length
of the string. If we look at the details, the dynamic behavior of this system may contain many
complexities: The mass of the object supported by the spring is in reality distributed over the object
rather than concentrated at a point. The stiffness of the spring is similarly spatially extended. The
spring has a finite mass and its response is unlikely to be exactly linear. However, if we disregard
these complexities and treat the system as a point mass supported by a massless spring, the behavior
of our model system and the true physical system turn out to be very similar. We have captured the
essential features of the system in our simple model, without including all the complex details that
have only small effects.
The process whereby we consider the mass to be concentrated at a point rather than distributed
in space is called lumping and we refer to the model as a lumped parameter model. Lumping is
a very powerful idea because it greatly simplifies the mathematical treatment of the system while
managing to retain the essentials. Masses and springs can both be lumped, i.e. replaced by a single
5

m
k

Figure 4: The simplest SDF system.


element. We will see later that another major component of our modelled systems, the damping,
can also be lumped. In all cases, the essential point to bear in mind is that we are modeling the real
physical situation. We know that the mass is not concentrated at a point in reality, but that we can
model it as being so and obtain a good approximation of the real situation.

2.2 Free vibration of a SDF system with no damping


We now start our mathematical analysis of the system shown in Figure 4. We will start off in the
normal way by drawing the free body diagram, deriving the equation of motion and finding solutions
to this equation. We will also look at the energy of the system and see how this too can be used to
derive an equation of motion. Of course, the two methods give the same results.

2.2.1 Equation of motion approach


This is the way we will tackle most problems in the course. First we draw the free body diagram as
shown in Figure 5.

m &x &
k x

Figure 5: The free body diagram for the system of Figure 4.


Now we can write down the equation of motion
mx + kx = 0 .

(4)

Exercise: 2
Consider the same system but with the effects of gravity included. Draw the free
body diagram and write down the equation of motion. Show that by appropriately
shifting the x coordinate one obtains the same equation of motion (4) as in the
gravity-free case. (Chose as x = 0 the static equilibrium configuration of the
system with gravity)
Before we try to solve this equation, we will classify it. The equation is second order in time because
the inertial term involves a second order derivative. The equation is linear, since the coefficients do
not depend on x, and it has constant coefficients because m and k do not depend on t. It is also
homogeneous because there is no term independent of x. In full then, the equation is a second order
homogeneous ordinary differential equation with constant coefficients.
To solve this equation, we need to find two complementary functions. We will not need a particular
integral because the equation is homogeneous. The equation is second order, so we expect to obtain
two complementary functions. The sum of these two functions is the general solution. We can see
by inspection that either of
x = sin(t) ,
7

x = cos(t) ,

(5)

is a solution of the equation. The constant must be chosen


p so that the m and k constants are
accounted for correctly. The value needed to do this is = k/m.
Exercise: 3
Verify the value given for . To do this, consider sin(t) first. Compute the
second derivative with respect to time and substitute into equation (4). From the
resulting equation you can calculate the relation between and k and m. Repeat
the calculation for the cosine solution.
From the complementary functions, we can now construct the general solution. It is
x = A sin(t) + B cos(t) .

(6)

It is easier to see what is going on if we rewrite the general solution for x in a different form. We
can write the solution as
x = C sin(t + ) .
(7)

2
2
where the constants are given by C = A + B and tan = B/A.
Exercise: 4
Show this is the case. Hint: start with x = C sin(t) and expand the sine using the
formula sin(X +Y ) = sin(X) cos(Y ) + sin(Y ) cos(X).
From this expression we can see more clearly what is happening. The constant C is the amplitude
of the vibration and is the phase. These two constants are undetermined in the general solution
and must be calculated from the initial conditions.
Lets now put some numbers into the formula, specify initial conditions, and see what the solution
looks like. Suppose m = 1 kg and k = 25 N/m. If we set the mass off at time zero at its equilibrium
position with a velocity of 1 cm/s, what does the subsequent motion look like? We calculate first
and find = 5 rad/s. The motion of the mass is therefore given by
x = C sin(5t/s + ) .

(8)

From the initial conditions, we know that x(0) = 0 which implies = 0. The velocity at any time t
can be found by differentiating the above equation to get
v(t) = x(t)
= [5/s]C cos(5t/s + )

(9)

and we now know = 0 so at time t = 0 the velocity is given by


v(0) = [5/s]C .

(10)

The initial conditions tell us that v(0) = 102 m/s so C = 2 103 m. The final solution then is
x(t) = [2 103 m] sin(5t/s) .
This solution is shown in Figure 6 for the first three seconds of motion.
8

(11)

displacement [mm]

-2
0

time[s]

Figure 6: The displacement of the mass in the system of Figure 4 as a function of time. The
parameters and initial conditions are given in the text below.
2.2.2 Energy approach
We can tackle the same problem discussed in Section 2.2.1 using an alternative approach based on
the energy of the system. We will find that we can actually derive the equation of motion once we
know the energy of the system, so all the results from the previous section can be re-obtained. The
advantage is that we dont need to consider the forces in the problem, which in some cases can
be difficult. On the other hand, the energy approach as discussed here works only if there is no
damping.
Consider the energy stored in the spring. The force exerted by a spring is F = kx where x is the
extension from the equilibrium length and k is the stiffness. The work done in extending the spring
by a length x is
Z
kx2
W = Fds =
.
(12)
2
The energy stored in the spring, its potential energy, is the negative of the work done on the spring
and we will call this U(x). We have
kx2
U(x) =
.
(13)
2

The kinetic energy of the mass we will call T . It is given by


mx2
,
2

(14)

mx2 + kx2
.
2

(15)

T=
so the total energy of the system E = U + T is
E=

The principle of conservation of energy tells us that E must be the same at all times, so dE/dt = 0.
We can use this information now to deduce the equation of motion. Differentiating the expression
for E with respect to time gives us
E = mxx + kxx
=0 .

(16)

x(m
x + kx) = 0.

(17)

We can factor this equation into


so either x = 0, which tells us that the system may be at rest, or mx + kx = 0 which is the equation
of motion derived above. The situation in which the system never moves is not of great interest to
us in a course on dynamics. The other result is very useful, however, because it shows us how to
obtain an equation of motion for a system once we know the total energy of the system. Note that
we have assumed that the energy in the system is conserved. In some cases, friction and other forces
dissipate energy to the surrounding. In these cases, the energy of the system decreases in the course
of time and so the approach we have used here cannot be applied.

2.3 Free vibration of a SDF system with damping


In reality, many systems that we meet have damping. The damping can be caused by friction in
sliding parts or by viscous effects in a fluid, for example. The mathematical specification of a
damping force can be quite tricky because of the different physical mechanisms that can cause
the damping. In all cases, however, the effects are similar: energy in the system is removed and
dissipated to the surroundings, often in the form of heat or noise. We will be primarily concerned
with viscous damping for which the mathematical specification is relatively simple. Just as with the
masses and springs, we will use lumped dampers to model the effects of damping in a system. This
means that although in reality the damping effects may be distributed throughout the system, in our
model the damping will occur at a well defined point.

2.3.1 Viscous damping and solution of the damped motion equation


The idea of viscous damping is that a system moving with a velocity v is slowed down by a force
proportional to v. Fast moving objects therefore encounter large forces, while slowly moving objects
10

encounter only small forces. The direction of the force is always to oppose the velocity so that the
system is always slowed by the damping, rather than speeded up. Mathematically, we can say that
the viscous damping force F is given by F = cv where c is a constant which determines the amount
of damping present in the system. Lets look at a simple system acting under the effects of viscous
damping.

m
c

Figure 7: A spring-mass-damper system.


Figure 7 shows how we represent symbolically a spring-mass system retarded by a viscous damper
(dashpot). Recalling that the force due to the viscous damper is cv = cx we can draw the free
body diagram for this system as shown in Figure 8.

m &x &
c x&
k x
m

Figure 8: The free body diagram for the system of Figure 7.

11

Notice the direction of the forces. The positive x direction is from left to right in the figure, so if
the mass is accelerating from left to right both x and x are directed left to right also. As we have
already discussed, the inertial force acts to oppose the acceleration we attempt to give the mass, so
acts right to left. Also, the damping force acts to oppose the motion, and since the mass moves from
left to right for positive x,
the damping force must act right to left. The equation of motion can now
be written down for this system. It is
mx + cx + kx = 0 .

(18)

This is a linear, homogeneous second order equation. The solutions of this equation are of the form
exp(t). We can see why this is the case by thinking about the properties needed for a function to
be a solution of equation (18). The coefficients in the equation are constants, so if we can choose
a function that under differentiation is equal to itself multiplied by a constant, then this is a good
candidate for a solution to (18). The function that satisfies this need is of course the exponential
function. The constant must now be chosen so that equation (18) is satisfied. To this end, we
substitute x = exp(t), x = exp(t) and x = 2 exp(t) into (18). We get
2 +

k
c
+ = 0 .
m
m

This quadratic equation (characteristic equation) has roots

c c2 4km
=
,
2m

(19)

(20)

and the corresponding solutions are given by


x+ (t) = exp(+t) ,
x (t) = exp(t) .
These are two complementary functions which solve equation (18). Since either of these functions
is a solution of (18), and the equation is linear, any combination Ax+ + Bx is also a solution (A and
B are arbitrary constants). The most general solution is therefore
x(t) = A exp(+t) + B exp(t) .

(21)

It is worth recognizing that the roots + and and therefore also the complementary functions
x+ (t) and x (t) are not always real. Of course, the final solution x(t) must be real, and this can be
ensured by taking, in a last step of the calculation, the real part of equation (21),
x(t) = Re[A exp(+t) + B exp(t)] .

(22)

The behaviour of this solution depends on whether the roots of the characteristic equation are
real or complex.

12

If c2 > 4km both roots given by equation (20) and, hence, also the complementary functions are
real. Inserting the values for + and into equation (21) gives
"
# !
"
# !
r
r
c
c
c 2 k
c 2 k
x(t) = A exp
+

t + B exp

t
.
(23)
2m
2m
m
2m
2m
m
This expression can be re-arranged and x(t) can written in terms of hyperbolic functions:
"
r
!
r
!#
c
c 2 k
c 2 k
x(t) = exp t C cosh
t + D sinh
t
2m
2m
m
2m
m

(24)

where C = (A + B)/2 and D = (A B)/2.


If on the other hand c2 < 4km then the roots are complex and the values of exp(+t) and exp(t)
will be complex quantities. The roots will be complex conjugates of each other, so the real part of
the roots is the same while the imaginary part is the same magnitude for each root, but of opposite
sign. We can therefore write
+ = R + iI ,
= R iI
where the real and imaginary parts R and I are
r
c
R =
2m

I =

k c 2

.
m
2m

As for the real roots case, the general solution is given by equation (21), but now the exponentials
and the constants A, B must be envisaged as complex numbers, A = AR + iAI and B = BR + iBI .
Inserting in Eq. (21) gives
x(t) = A exp(+t) + B exp(t)
= exp(Rt) [(AR + iAI ) exp(iIt) + (BR + iBI ) exp(iIt)]
= exp(Rt) [(AR + BR ) cos(It) (AI BI ) sin(It)
+i((AI + BI ) cos(It) + (AI BI ) sin(I t))]
"
r
!
r
!#
c
k c 2
k c 2
= exp( t) E cos

t + F sin

t
2m
m
2m
m
2m

(25)

where in the last line I have taken the real part and used new constants F = AR + BR and G =
AI + BI to replace the old ones. Note the similarity to equation (24).
The form of the solutions (23) and (25) is very interesting. Equation (23) is the sum of two exponential decays. This means that the motion of a system for which this equation is applicable, i.e.
when c2 > 4km, is just a decay towards zero. No oscillations can be seen, and we refer to the system
13

Figure 10: An example of the solution for an


under-damped system. The envelope of the oscillations is a decaying exponential.

Figure 9: An example of the solution for an


over-damped system. The solution exhibits no
oscillations.

as being over-damped. On the other hand, when c2 < 4km, equation (25) shows that the motion
consists of an oscillatory part given by the sinusoidal functions in the brackets, multiplied by an exponential decay. The system therefore oscillates, but the amplitude of the oscillations decays. The
system is referred to as being under-damped. Figures 9 and 10 show plots of what these solutions
look like.
The two different types of behaviour, over-damped and under-damped oscillation, can be identified
using the ratio c2 /4km. For c2 /4km > 1 the system is over-damped and for c2 /4km < 1 the system
is under-damped. For c2 /4km = 1 the system is said to be critically damped. We wont look at this
case because in practice the ratio is never precisely equal to one.

2.3.2 Canonical form of the damped free vibration equation


We have seen how to solve the damped free vibration equation for a mass-spring damper system.
We considered the equation
mx + cx + kx = 0 .
(26)
When we are dealing with other vibrating systems, the coefficients in this differential equation may
change, and even their physical meaning may be different. We will see an example of this in the
next section. However, the basic structure of the equation of motion and the method of solution
are always the same. Rather than finding the solution separately for each case, we write equations
of this type in a general form. This is usually referred to as canonical form. We can then use the
solution to the canonical equation straight away, rather than having to redo all the algebra. To this
end, we use two fundamental parameters. The first parameter is the frequency of vibration of the
system without damping. We also refer to this as the natural frequency 0 of the system. For the

14

mass-spring system described by equation (26), we find that the natural frequency is
r
k
.
0 =
m

(27)

To characterize the damping, we introduce the damping ratio which determines whether the system is under-damped or over-damped. For our mass-spring-damper system the damping ratio is
defined by
c2
2 =
.
(28)
4km
As we have seen in the preceding section, > 1 gives over-damping, and < 1 gives under-damping.
We can rearrange equation (28) to give a useful alternative expression for the damping ratio:
c
c
.
=
=
2 km 2m0

(29)

We can now write equation (26) in terms of and 0 instead of k, m and c. Dividing through by m
we have
k
c
(30)
x + x + = 0 .
m
m
and from the relations for 0 and
k
= 20
m

c
= 20 ,
m

(31)

we find that this can be written as


x + 20 x + 20 x = 0 .

(32)

By writing our equations in this canonical form, we can see at once the value of the damping
parameter and the undamped natural frequency. The true frequency of vibration is the coefficient of
t in the sine and cosine terms in equation (25). It is called the damped frequency d and is given by
r
k c 2
d =

(33)
m
2m
We can express this in terms of 0 and :
d = 0

p
1 2 .

(34)

This gives us the result that if damping is very small then d 0 . We can also write the solution,
equation (25), using these new parameters:
x(t) = exp(0 t)[E cos(dt) + F sin(dt)] .

(35)

Just as we did with the undamped system, we can rewrite the sum of the sine and cosine term as a
sine term with a phase shift. We get an equivalent formula to the above:
x(t) = H exp(0 t) sin(dt + ) .
15

(36)

The constants H and can be determined from the initial conditions.


Another parameter which is often used to characterize damped vibration is the logarithmic decrement of the oscillation. The logarithmic decrement is defined as the natural logarithm of the ratio
of any two successive maxima of the oscillation. It is related to the damping ratio by
2
=
1 2

(37)

Exercise: 5
Use the solution (36) to derive this expression from the ratio between two successive maxima of the oscillating solution.
Despite all the mathematics above, we can see that the effect of damping on the vibrating system
is actually quite straightforward. If the damping is larger that a critical amount, given by = 1,
the system is over-damped and simply relaxes to its equilibrium position without any oscillation.
The more interesting situation from our point of view is when the damping is less than the critical
damping. In this case, the addition of the damping alters the frequency of the oscillations, and
makes their amplitude decay gradually to zero.
To use the canonical form to best advantage, we first write down the equation of motion for a system
using the free body diagram. We can then identify the parameters 0 and , and from these deduce
whether the system is under or over-damped, the frequency of vibration and the rate of decay of the
oscillations (if present).

2.3.3 Rotational vibration


Until now we have considered systems where masses move along a given fixed direction, i.e. the
motion is translational. However, in many important cases vibration is related to the rotation of
parts. In this case our procedure of setting up the equations of motion must be modified. Figure
11 shows a system which may exhibit rotational vibrations: A square block of mass m can rotate
around the axis A and is attached to the walls by a spring of constant k and a damper with damping
constant c. The spring constant is such that the block is at rest in the position shown.
Lets see what happens when we rotate the block to the right by a small angle . The top right corner
of the block (where the damper is attached) moves to the right by l sin and the bottom right corner
where the spring is attached moves downward by the same amount. This leads to a compression of
the spring and, if the rotation occurs at finite speed, to a viscous damping force from the damper.
To work out the spring force and the damping force, we use linearized relations by noting that, for
small angles , sin . Hence the spring and the damper are compressed by x = l. The velocity

of compression of the damper is obtained by taking the time derivative, x = l .


The corresponding forces are drawn in Figure 12. Now we have to keep in mind that the block
16

c
m

k
l
Figure 11: A system undergoing rotational vibration.

rotates around the axis A, so we have to consider the sum of all moments with respect to this axis.
and their moments with respect to A are kl 2 and
The spring and damper forces are kl and cl ,
2
In addition we have to consider the effect of inertia. This leads to a moment IA where IA is
cl .
the moment of inertia of the block with respect to the axis A and = is the angular acceleration.
Exercise: 6
Calculate the moment of inertia of the block around the axis A. Assume that the
block is homogeneous. (Result: IA = 2ml 2 /3)
The sum of the moments around A must be zero. Hence we end up with the equation of motion
IA + cl 2 + kl 2 = 0 .

(38)

or, after inserting IA = 2ml 2 /3 and dividing by 2l 2 /3


m + (3c/2) + (3k/2) = 0 .

(39)

This equation is very similar to the equation of motion for the translational motion of a mass-spring17

c lf &

c l 2f&

k l 2f
J f &&

k lf

Figure 12: (Left) forces and (right) moments acting in the system of Fig. 11
damper system, equation (18), and can be written in the same canonical form. To this end, we divide
by the pre-factor of the term with the highest derivative to obtain
3c 3k
=0 .
+
+
2m
2m

(40)

We identify this with the general canonical form as given by equation (32) - here we have the angle
instead of x as the dependent variable, but everything else remains the same:
+ 20 + 20 = 0 .

(41)

By comparing coefficients, we find


20 =

3k
2m

20 =

3c
,
2m

and, hence, the natural frequency and damping ratio for this system are given by
r
r
3k
3c2
0 =
, =
.
2m
8km

(42)

(43)

We can now use the general solution of the canonical vibration equation by inserting these parameters into equation (35) or (36) and determining the remaining unknown parameters from initial
conditions.

18

x
c
m

f(t)

k
Figure 13: An example of a simple system subjected to forcing

2.4 Forced vibration


We will now consider the vibration of a system which involves an external force. This will we
call the forcing. We will consider both periodic and non-periodic forcing. We will make use of
the complex exponential method once again. You may have found it possible to work through
the problems so far without using complex exponentials. You will struggle to do this for forced
vibration. Please make sure you are familiar with the workings of the complex exponential method,
and make sure you can do the questions Ive set on it. Once you are happy with the method, youll
probably find that the material that follows is not too complicated.

2.4.1 Periodic forcing


Consider the system shown in Figure 13. The forcing f (t) is quite general; we will first consider
the case where f (t) is a periodic function, so f (t) = f0 cos(t).
Exercise: 8
Draw the free body diagram for Figure 13 and deduce the equation of motion.
The equation of motion for the system shown in Figure 13 is
mx + cx + kx = f (t) = f0 cos(t) .

(44)

Before we solve this equation we consider the simplest case where the applied force is a constant
( = 0). A constant force leads to a constant elongation of the spring by x0 = f0 /k. We call x0 the
static response of the system.

19

Now we put the equation into canonical form. We first divide by m to get
f0
cos(t) .
(45)
m
The right-hand side of this equation can be written in terms of the static response and the natural
frequency:
x + 20 x + 20 x = 20 x0 cos(t) .
(46)
x + 20 x + 20 x =

This is the form which we will always use in the following. Looking at the problem from a mathematical point of view, I can see that the equation of motion is an inhomogeneous equation and so
the solution will be given by the sum of two complementary functions and a particular integral. The
complementary functions are the solutions of the homogeneous problem and have been discussed in
the previous section, they will both be damped oscillations, so after a reasonable time they will have
decayed to small values. At this stage, Im interested in how the system responds after quite some
time so that these initial effects will have died away. Thats what we call the steady-state response
of the system. I will therefore ignore the complementary functions in what follows.
Now let us take a closer look at equation (46). The left-hand side (LHS) basically involves x multiplied by various constants and differentiated. The solutions to this sort of equation are things like
x exp(t), where may be complex. The right-hand side (RHS) is a cosine, but I can write that
also in complex exponential form
x + 20 x + 20 x = 20 x0 exp(it)

(47)

and regard x as a complex variable. I can recover the original equation of motion by looking at just
the real part of this equation. Since the equation is linear, I can say that if x = xR + ixI is a solution
then xR is a solution to the real part of the equation and xI is a solution to the imaginary part. I can
therefore go ahead and solve for the complex variable x and take the real part of x at the end of the
calculation. This real part will be a solution to the original equation of motion.
I expect the steady-state response to be
x(t) = X exp(it) ,

(48)

where x and X are complex. Substituting this expression gives


X=

x0 20
.
2 + 2i0 + 20

(49)

To bring this complex amplitude into a more useful form, I use that any complex number A + iB can
also be written in the form of a complex exponential:
A + iB = r exp(i)
where r2 = A2 + B2 and tan = B/A. When we write X in equation (49) in this form, we get
x0
X = q
exp(i) ,
2
1 2 /20 + 42 2 /20
20

(50)

(51)

where

2/0
= arctan
.
1 2 /20

(52)

We now insert insert X into the solution (48):


x0
x(t) = q
exp(i[t + ]) ,
2
2
2
2
2
2
1 /0 + 4 /0

(53)

As I explained above, this is the solution to the complex form of the equation of motion. The
physical solution is simply the real part of it. I use that

and find

cos(t) = Re[exp(it)] .

(54)

x0
xR = q
cos(t + ) .
2
2
2
2
2
2
1 /0 + 4 /0

(55)

What does this mean? There are several interesting points to extract from this solution. Firstly, it is
basically another sinusoidal function of the form
x = A cos(t + ) ,

(56)

x0
A = q
2
1 2 /20 + 42 2 /20

(57)

where the amplitude of the vibration is

and the frequency is the same as that of the forcing. However, there is a phase shift between the
forcing and the response. The system therefore responds to the forcing by vibrating at the same
frequency but with a delay,

2/0
= arctan
.
(58)
1 2 /20
The phase lag depends on the frequency of forcing. For small , is small. For = 0 ,
= /2, and for , .
Exercise: 9
What does a phase shift of radians mean physically?
Exercise: 10
Write down the values of for = 0.1 and = 0, 0 + , 0 , where is a
very small number. Verify that the values of the phase shift for small are 0,
for = 0 , = /2, and for , .
When you use your pocket calculator to get the phase shift , you must be careful when taking the
inverse tangent of a negative quantity. The inverse tangent is only defined up to a factor of n where
n is an integer number, and you may have to subtract to get the phase shift right.
21

We now look how the amplitude behaves as a function of . Take the expression for A,
x0
A = q
2
1 2 /20 + 42 2 /20

(59)

and consider the case when the damping ratio is small. Now, the squared terms inside the square
root are always positive, so since they appear in the denominator, the amplitude is largest when they
are small. The first term is zero when = 0 . So when the damping is small and the forcing is
the same frequency as the frequency of free vibration, the amplitude is a maximum. This is called
resonance. For small values of the damping ratio we can often use the approximate expression
x0
A=
.
(60)
|1 2 /20 |
This expression shows that the resonance peak is located at approximately 0 (this follows from
the small damping ratio assumed in equation (59)). The amplitude of the dynamic response at
resonance for small is 1/(2).
Exercise: 11
Calculate the frequency and amplitude of resonance (the frequency and amplitude
where the response has its maximum) for the case where is not small.
Most of what weve been dealing with here can be neatly expressed in two graphs; one for the amplitude and one for the phase as functions of the frequency of forcing. The graph for the amplitude
is probably the most useful form of the frequency-response graph. It tells us a great deal about the
dynamics of a system, all on a single plot. If we plot on (x, y) axes, then we let
y = A/x0 ,

x = /0 ,

(61)

and so from equation (59) we have


1
y= p
2
(1 x )2 + 42 x2

(62)

which determines the shape of the frequency-response graph. Because at resonance the response is
usually very large, you will often find these graphs plotted in logarithmic coordinates (i.e. plot ln y
against ln x).
Exercise: 12
Use MATLAB to deduce the shape of these graphs for a number of values of and
add sketches to your notes. Deduce the height and width of the resonance peak and
label it.
2.4.2 Force transmission and vibration damping
We now consider the following question: Given the amplitude f0 of a periodic forcing acting on the
mass m in Figure 13, what is the amplitude of the force transmitted to the support?
22

Adding the spring and damper forces, we see that the transmitted force is given by
fs (t) = kx(t) + cx(t)

(63)

Inserting the solution given by equation (56), we find


fs (t) = kA cos(t + ) cA sin(t + )

(64)

Now we use that this can be written as


fs (t) = fs,0 cos(t + 0 )
where fs,0 =

k2 A2 + c2 2 A2 . Hence the amplitude of the transmitted force is

p
x0 k2 + 2 c2
2
2
2
fs (t) = A k + c = q
2
1 2 /20 + 42 2 /20

and using the relations x0 = f0 /k, c2 = 4km2 and 20 = k/m this can be written as
v
u
u
1 + 42 2 /20
fs (t) = f0 t
= f0 TR
2
1 2 /20 + 42 2 /20

(65)

(66)

(67)

The ratio between the force acting on the mass and the force transmitted to the surroundings is
called the transmissibility ratio
v
u
u
1 + 42 2 /20
.
(68)
TR = t
2
1 2 /20 + 42 2 /20
Exercise: 13
Sketch the curves of TR plotted against x = /0 for x = 0 . . . 5 in intervals of 0.5.
Use values of = 0.1, 0.2, 0.5 and compare the different dampings. Plot the same
graph on logarithmic scales.
The form of the transmissibility ratio shows that at the resonant frequency, large values of damping
are good because they reduce force transmission. However, at high frequencies the opposite is the
case and the most efficient way to reduce force transmission is to make 0 as small as possible (soft
springs, large masses). In practical circumstances it is vital to be aware of this compromise.

2.4.3 Moving boundary condition and amplitude transmission


Until now we have considered forced vibration where an external force is acting directly on a vibrating mass. A different type of excitation is shown in Figure 14.
23

x
c
m

k
z (t)
Figure 14: A system with moving boundary excitation
We assume that the support of the system moves periodically according to z(t) = z0 cos(t) and ask
for the amplitude of the steady-state response of the system.
The equation of motion is
mx + c(x z) + k(x z) = 0.

(69)

We collect the terms involving z on the right-hand side and bring the equation into canonical form
by dividing through m
x + 20 x + 20 x = 20 z0 sin(t) + 20 z0 cos(t).

(70)

We now bring also the right-hand side into canonical form. To this end, we write
20 z0 sin(t) + 20 z0 cos(t)
= 20 z0 [cos(t) (2/0 ) sin(t)]
q
= 20 [z0 1 + 42 2 /20 ] sin(t + 0 ) .

(71)

We finally shift the time axis to get rid of the phase 0 (we are interested in the long-time response
so the choice of the start time really doesnt matter) and find that the equation of motion assumes
the canonical form
x + 20 x + 20 x = 20 x0 cos(t)
(72)
q
with x0 = z0 1 + 42 2 /20 . We can now use the solution (56), x(t) = A cos(t + ) with the
amplitude A given by (59). As a result we find that the ratio of the vibration amplitudes of the mass
and of the support is given by
A/z0 = TR
(73)

24

with the same transmissibility TR as in the previous section. Hence, equation (68) gives us both
force and amplitude transmission, and what has been said about the influence of the parameters 0
and on vibration isolation also applies to the present case.

2.4.4 Rotating out of balance rotor

w t
e

G
O
m

x (t)
M

Figure 15: Schematic diagram of a machine with an out of balance rotor. The centre of mass of the
rotor is located at G.
Many machines have rotating parts which can vibrate. The system we will look at are the vibrations
of a machine which are caused by a rotor which is out of balance. Figure 15 shows a machine of
mass M which contains a rotor of mass m with an eccentricity of e. The machine is mounted on
a solid floor by a spring of stiffness k and a damper with damping constant c and can move in the
vertical (x) direction only.
From the free-body diagram shown in Figure 16, we find that the equation of motion for the machine
and rotor in the x direction is
(m + M)x + cx + kx = me2 cos(t) .

(74)

We follow our usual procedure of casting this into canonical form,


x + 20 x + 20 x = 20 x0 cos(t) .
where now

r
0 =

s
k
,
m+M

c2
,
4k(m + M)
25

x0 =

(75)
me 2
.
m + M 20

(76)

m ew
2

m x &&
c x&

x (t)

M x &&

k x

Figure 16: Free body diagram for the system of Figure 15.
It follows from the results of Section 2.4.1 that the steady-state solution is given by x(t) = X cos(t +
) where the amplitude X is given by
X=

2 /20
me
q
m + M (1 2 /2 )2 + 42 2 /2
0

(77)

and the phase is given by


tan =

2/0
.
1 2 /20

(78)

If we look at the amplitude and the phase of the motion for low, resonant and high frequencies we
see that
0

X 0

= 0

X = me/[2(m + M)]

X me/(m + M)

0 ,
= /2 ,
.

(79)

The highest possible amplitude occurs at resonance in a system for which m M. In this case, the
maximum amplitude is X = e/. If the damping is small this can be very large.
Exercise: 14
Use the results of the previous sections to determine the amplitude of the force
which is transmitted to the ground. What is the force transmitted at resonance?

26

2.4.5 Non-periodic forcing


In certain cases we wish to analyse the response of a system to non-periodic forcing. One example
of this might be a sudden shock. We will consider a shock first of all, and then show that more
complicated forcings can be considered as a amalgamation of shocks one after the other. This will
give us a formula with which we can calculate the response to an arbitrary forcing.
Consider a spring-damper system such as that shown above in Figure 13 two pages above, with the
forcing function shown in Figure 17.

F o rc e

~ 1 /e

T im e

Figure 17: A short-lasting force.


The force shown contains a parameter which we will take to be small. This means that the force
lasts only briefly, but is large - it can be envisaged as a heavy kick.
We define the impulse to be the integral
I=

f (t)dt .

(80)

What is the response of the system to an impulsive shock like this? Lets be clear about the problem
we wish to solve. Given the spring-damper system of Figure 13 subjected to an impulse of size I,
what is its response? We will assume that initially the system is in its equilibrium state.
Its actually quite straight-forward to work out. We will find that the effect of the forcing is just to
give the system an initial kick, and from then on it will behave like a freely vibrating system. If we
27

can work out the size of this kick, this will allow us to set the initial condition for the subsequent
freely vibrating behaviour.
All we need to do is to recognise that Newtons second law can be written as
Fm =

dp
dt

(81)

where p = mv is the momentum, and Fm is the force acting on the mass m. Therefore, the momentum
of the mass is given by
Z
t

p(t) p(0) =
0

Fm dt .

(82)

When the initial kick is very strong and very short, the effect of the springs and dampers during the
kick can be ignored because the external force is much larger than the spring and damper forces.
Initially, v(0) = 0 because the system starts at rest, so p(0) = 0 and the momentum of the mass m
just after the shock is given by
p(t = ) = mv() = I ,
(83)
and so
v(t = ) = I/m .

(84)

Now, if is very small (the kick is very short), we can consider the motion of the system to start
at time 0 with the velocity v = I/m, and since the forcing is zero after time we need only consider
the motion of the freely vibrating system. We found above that the solution for a freely vibrating
system is
x(t) = exp(0 t)[A cos(dt) + B sin(dt)]
(85)
and if we set the initial condition x(0) = 0, it follows that A = 0, i.e.,
x(t) = B exp(0 t) sin(dt) .

(86)

The other initial condition is x(0)


= I/m. We find:
x(t)
= B exp(0 t)[0 sin(dt) + d cos(dt)] ,

(87)

x(0)
= Bd = I/m ,

(88)

so
hence
B=

I
,
md

(89)

and the solution we require is


x(t) =

I
exp(0 t) sin(dt) .
md

28

(90)

This expression gives us the response of a system to an impulse of size I delivered at time t = 0. It
is usual to define the unit impulse response function h(t) to be
h(t) =

1
exp(0 t) sin(dt) ,
md

(91)

and so the response to an impulse I is


x(t) = Ih(t) .

(92)

We often find that we can approximate the behaviour of other forces as shocks, even if they last for
quite a long time. The spring-mass system has a time scale for its vibrations of T0 = 2/0 and if
the time over which a force acts is much smaller than T0 we can usually treat it as a shock and write
down the response using the unit impulse response function h(t).
Exercise: 17
Sketch the behaviour of a mass-spring-damper system after an impulse is applied
For complicated non-periodic forcing, we can generalise the above method. Consider the forcing
function shown in Figure 18:

F o rc e

t-u

u u + d u

T im e

Figure 18: A general non-periodic forcing.


We can imagine that the forcing F(u) is split up into lots of pieces defined over intervals of time du
(note that for clarity Im using the variable u to denote time rather than t - the reason will become
clear later). Each of these can be thought of as a little impulsive shock. If we consider the response
at time t, we can think of this response as being built up of many responses to these shocks remember, we are allowed to build up solutions like this because the equation is linear and so the
super-position principle holds.

29

Lets consider the response at time t due to the shock shown at time u. The impulse due to this
shock is
I = F(u)du
(93)
At time s after the application of a shock of impulse I, we know that the response due to the shock
is
x = h(s)I
(94)
Now, the time after the shock is shown by the arrows and is t u so the response at time t due to the
shock at time u is
xu = h(t u)F(u)du ,
(95)
where Ive used the subscript u to denote that this is the part of the response due to the shock at time
u. The full response is the total of all the shocks from time 0 to time t, i.e.,
x(t) =

Z t
0

xu du =

Z t

F(u)h(t u)du ,

(96)

and so we can calculate the response using this expression, provided we can do the integral. In full,
the formula for the response is
x(t) =

Z t
F(u)
0

md

exp(0 (t u)) sin(d (t u))du

(97)

which is called the convolution formula. For the case of small or no damping, this expression
simplifies a bit to
Z t
F(u)
x(t) =
sin(0 (t u))du
(98)
0 m0
which is in general easier to integrate. This I will refer to as the convolution formula without
damping.

2.4.6 Shock damping and shock response spectrum


To assess whether a vibrating system is capable of effectively damping shocks, we consider the
shock response spectrum. This is a plot relating the response to the forcing duration, but plotted
in a clever choice of variables: On the y-axis, the ratio of the peak response x and the peak static
response x0 is plotted. The peak response is the highest maximum of the response to the shock,
which is also sometimes called the maximax. The peak static response is just the response that
would be caused by a force equal to the peak value of the forcing applied statically. On the x-axis,
the frequency of free vibration divided by the frequency scale of the forcing is plotted. For a force
of duration , the frequency scale is simply 2/. Therefore, if the peak value of the shock force is
F0 and the system has a mass m, the peak static response is
F0 /k = F0 /(m20 ) .
30

(99)

Peak amplitude [F0/(m 0 )]

0
0.0

0.5

1.0

1.5

2.0

Shock duration [2/ 0]

Figure 19: Shock response spectrum for a rectangular pulse. Full line: Primary SRS; Dashed line:
Residual SRS
If the shock force lasts for a time then the frequency of free vibration of the system divided by
the frequency scale of the forcing is 0 /(2). On the shock response spectrum, we plot xm
20 /F0
against 0 /(2) .
The shock response spectrum depends in general on the shape of the force pulse. As an example,
we consider the response of an undamped mass-spring system to a rectangular pulse,

F0 , t
F(u) =
(100)
0 else .
The response to this pulse is obtained from equation (98). We consider separately the two cases
t and t > :
(i) For t the response is
x(t) =

F0
m0

Z t
0

sin(0 (t u))du =

F0
[1 cos(0t)] .
m20

(101)

The maximum of the response is at tmax = (2nmax 1)/0 with the peak value x = 2F0 /m20 . nmax
is a positive integer number. Hence, this maximum is reached only if > /0 .

31

(ii) For t > we find


F0
x(t) =
m0

Z
0

sin(0 (t u))du =

F0
[cos(0 (t )) cos(0t)] .
m20

(102)

The maxima are in this case given by


x =

2F0
| sin(0 /2)| .
m20

(103)

The shock response spectrum (SRS) is plotted in Figure 19. The primary SRS shows the absolute
maximum of the response which occurs during the time of the shock, and the residual SRS shows
the absolute maximum of the response occurring after the end of the shock. It is interesting to note
that for shock durations which are integer multiples of the period of vibration, there is no residual
response at all.
We now determine the conditions under which the maximum of the shock response falls below the
static response, i.e. where we have shock damping. This requires that
x = sin(0 /2) < 1/2 .

(104)

Hence, shock damping is achieved if the pulse is sufficiently short such that 0 < /3, i.e. the
ratio between the shock duration and the period of free vibration must be /T < 1/6. We may
compare this with the criterion for vibration damping which follows
from equation (57), where we

have observed that vibration damping is achieved if 0 / < 2, i.e.,


the ratio between the period
of the forcing and the period of free vibration must be less than 1/ 2. In either case and efficient
strategy to achieve damping is to make the natural frequency of the system as small as possible
(large masses and/or soft springs).

32

3 The theory of systems with many degrees of freedom


The generalisation of what we have done for one degree of freedom systems to many degrees of
freedom is straight-forward provided we do it in the right way. What we will find is that if we use
matrices to represent our equations, then the free body diagram equation of motion complex
exponential method we have used works just as well for the many degree of freedom systems. The
main complication is the maths, which can get rather lengthy. In fact, we will often have to resort
to computers and numerical methods in order to solve the equations. We will also find there are one
or two extra things going on in these systems that simply do not occur in a single degree of freedom
system.

3.1 Free vibration


We will start our study of multiple degree of freedom systems with a relatively simple example
shown in Figure 20.

x
k
m

x
1

2 m

Figure 20: A two degree of freedom system consisting of two masses connected by springs.
Exercise: 15
We have discussed in the Introduction how to obtain equations of motion for this
kind of system. Draw a free body diagram and write down the equation of motion
for each mass.
If we write down the equations of motion for the two masses in the system shown and collect similar
terms together we get

mx1 + 2kx1 kx2 = 0 ,


33

2mx2 kx1 + 2kx2 = 0 .

(105)

I can represent these equations of motion in matrix form:

m 0
0 2m

x1
x2

2k k
k 2k

x1
x2

=0 .

(106)

Exercise: 16
Multiply out the matrix equation above to check that you do recover the two equations of motion.
I could write this matrix equation symbolically as
M~x + K~x = 0 ,
where bold quantities represent matrices and the
and the stiffness matrices. They are given by

m
M=
0

2k
K=
k

(107)

arrows represent vectors. M and K are the mass


0
2m
k
2k

(108)

Written in this form, you can see that the equation looks really similar to the single degree of
freedom equations we are used to solving. I can solve the equation using the complex exponential
method. I write
~x = ~A exp(it) ,
(109)
where the usual x and A quantities I get in the single degree of freedom systems have become
vectors. Alternatively, I can write out this equation in long hand as the pair
x1 = A1 exp(it) ,
x2 = A2 exp(it) .

(110)

I calculate the derivatives and substitute into the equation of motion in the usual way to get
2 M~A exp(it) + K~A exp(it) = 0

(111)

Cancelling out the exponentials and collecting terms gives me

which I can write as

(2 M + K)~A = 0

(112)

B~A = 0

(113)

34

with
B = 2 M + K

(114)

Now a general result in the theory of linear algebra that you may be aware of is that for the system
of equations B~A = 0 to have nontrivial solutions, it must be true that the determinant of B must be
zero,
|B| = 0 .
(115)
If this is not familiar then take the time to go over your maths notes to refresh you memory. Replacing B, we have deduced at this point in the calculation that
| 2 M + K| = 0
If I now put in the expressions for the matrices I get the determinant

2k m2 k

=0 .
2
k
2k 2m

(116)

(117)

Now I can expand this determinant out to get


(2k m2 )(2k 2m2 ) k2 = 0

(118)

2m2 4 6km2 + 3k2 = 0 ,

(119)

which rearranges to
which is a quadratic equation in 2 . I can solve the quadratic to find two roots. I get
v
u
!
uk 3
3
= t

.
m 2
2

(120)

Negative frequencies dont really mean anything. I therefore have two solutions which I will label
+ and . They are given by
v
u
!
uk 3
3
+ = t
+
,
(121)
m 2
2
v
u
!
uk 3
3
t

.
(122)
=
m 2
2
This result is important. I have managed to calculate the frequencies of free vibration of the system
of two masses. Interestingly, there are two possible frequencies for free vibration. Either can occur
i.e. the system will resonate at either of these frequencies. It turns out, as we shall see, that a system
with n degrees of freedom has n frequencies of free vibration. It is usual to refer to vibration at these
frequencies as modes of vibration or vibration modes.
Now I have evaluated , I can go back to equation (100) and put in either + or . For each value
I can use the matrix equation to calculate the ratio A1 /A2 .
35

Exercise: 17
Use the values I have obtained for + and to write out equation (112) in full,
using the definitions of M and K in equation (108). Expand out the matrix equation
into two ordinary equations, and solve to find the ratio A1 /A2 .
For the frequency + , the ratio A1 /A2 is -2.73 and for , the ratio A1 /A2 is 0.73. What does all
this mean? Lets go back to the single degree of freedom system, and consider what happened there.
We made a substitution of the form
x = A exp(it)
(123)
in pretty much the same way as we have here, calculated the allowed value of , and that was it.
We couldnt determine the value of A, the amplitude of vibration, without more information. This
extra information came in the form of the initial conditions. For this two degrees of freedom system,
we have found two possible values of , and for each a value of the ratio A1 /A2 . Now, from the
substitution we made
~x = ~A exp(it) ,
(124)
we can see that the first component A1 of the vector ~A represents the amplitude of vibration of
the left mass since it is measured by coordinate x1 , and the second component A2 represents the
amplitude of vibration of the right mass. Once again, we cannot deduce the values of both A1 and
A2 without knowing the initial conditions. However, we can deduce the ratio of the two, and this is
interesting. For the first mode of frequency + , we know A1 /A2 = -2.73. In this case, the amplitude
of the left mass is more than twice that of the right. The minus sign indicates that the direction of
oscillation of the left mass is opposite to that of the right, i.e. they are in antiphase. For the second
mode, A1 /A2 =0.73, and so both masses are in phase, with the amplitude of vibration of the right
mass a little larger than that of the left one.
We can write for = +

~A = c+

and for =

2.73
1

~A = c

0.73
1

(125)

(126)

where we have introduced constants c to reflect the fact that the amplitude is not known. Each of

2.73
~x(t) = Re c+
exp(i+t)
(127)
1
and

0.73
~x(t) = Re c
exp(it)
1

(128)

are valid solutions of the equation of motion. The constants c+ and c are in general complex, so
by taking the real part of the complex exponential we get a combination of sines and cosines, or
alternatively a sine with a phase shift.
36

The general solution can be written as a sum of both solutions,

2.73
0.73
x(t) = c+
sin(+t + + ) + c
sin(t + ) ,
1
1

(129)

where the mode amplitudes c+ and c are now real and together with the phases + and must be
determined from initial conditions. So we need four initial conditions. In general, for a system with
n degrees of freedom 2n initial conditions are required.

3.2 Eigenvalue problems


We found in the above section that we could write our equations of motion in matrix form as
M~x + K~x = 0 ,

(130)

and indeed this is always the case, even for many more degrees of freedom. The matrices are square,
and usually symmetric although there are occasions when they are not. Using the substitution

we can transform the equation to


or rearranging terms

~x = ~A exp(it) ,

(131)

2 M~A + K~A = 0

(132)

(M1 K)~A = 2~A .

(133)

Now this last form is called the matrix eigenvalue problem, and it arises from many different physical situation. Because it occurs so often, it has been studied very thoroughly. Moreover, there are
many computer packages that will quickly solve it for us. Lets consider what the solutions are like
first, then see how a computer can be used to help in the calculations.
If we write
and

Y = M1 K

(134)

= then the problem we wish to study is


Y~A = ~A

(135)

The mathematics tells us that if Y is a n n matrix then there are n values of that allow this
equation to have non-trivial solutions, and each of these has associated with it a vector ~A. The
values of are called eigenvalues and the associated vectors are eigenvectors.
We have seen that we can take our vibration problem, cast it in matrix form and then manipulate the
matrices so that it is a eigenvalue problem. Using a package such as MATLAB, the solution of this
eigenvalue problem is very easy. In this way, we can tackle systems with quite a large number of
degrees of freedom that would be very hard indeed if we did not have access to a computer.
37

3.3 Forced vibration


Just as with the single degree of freedom systems, we would like to be able to calculate what happens
with a forced system. The path from the free body diagram to the equations of motion is no more
complicated than we have met already. Once again, we can write the equations in matrix form so
we end up with something like
M~x + K~x = ~F cos(t) ,
(136)
where the new part is the RHS consisting of a forcing term proportional to cos(t). Many times
only one component of ~F will be non-zero, but that need not concern us now. Using what we know
about matrices, solving this really quite difficult problem is surprisingly easy. Just as with the single
degree of freedom systems, we introduce a guessed solution of the form
~x = ~X cos(t)

(137)

where you should note that we have been able to use cos(t) directly rather than the complex
exponential because there is no damping and so we know that the response is either in phase or in
antiphase with the forcing. Substituting this into the equation of motion gives
2 M~X cos(t) + K~X cos(t) = ~F cos(t) .

(138)

and we can cancel the cosines from each side so we have an equation of the form
Z~X = ~F

(139)

Z = 2 M + K .

(140)

where
The matrix equation (139) corresponds to system of equations for the components of the vector ~X.
For small matrices (systems with only a few degrees of freedom) this can be solved by hand. For
larger systems, one uses a formal solution which can be computed numerically. To this end, we
have simply to determine the inverse matrix of Z:
~X = Z1 ~F .

(141)

The inverse of a matrix can be calculated using MATLAB. Hence, we may simply use MATLAB to
calculate the inverse matrix Z1 and multiply this with the vector ~F which contains the amplitudes
of the external forces.
Exercise: 18
Go back to the two spring system we defined above and add a forcing term of the
form F0 cos(t) to the left mass. Use the above method to deduce an expression
for the frequency response curve.

38

3.4 Orthogonal modes


I want to conclude our consideration of the theory of multiple degree of freedom systems with some
more matrix algebra. Make sure your understanding of matrix algebra is up to scratch before we
start. The following considerations are rather formal, but they lead to a surprising result: A vibrating
system with n degrees of freedom can always be transformed in such a manner that it behaves like
n independent single degree of freedom systems.
The matrix equation of motion for a n degree of freedom system is
M~x + K~x = 0 .

(142)

If we introduce the vector of coordinates ~A such that

then we have

~x = ~A exp(it) ,

(143)

K~Ai = 2i M~Ai ,

(144)

where we know that there should be n possible values of and Ive acknowledged this in the
equation by labelling these different possible s and the corresponding ~A s with an index i which
can run from 1 up to n. Now Im going to prove an important result concerning the vectors . Firstly
I will multiply the above equation by the transpose vector ~ATj to give me
~ATj K~Ai = 2i ~ATj M~Ai .

(145)

It will become clear later just why I want to do this. For the moment, just make sure you understand
what this means - write out example matrices for yourself if you like. I will also make use of exactly
the same equation written out with the i and j subscripts interchanged,
~ATi K~A j = 2j ~ATi M~A j .

(146)

Now, the theory of matrix algebra shows that


(AB)T = BT AT .

(147)

Using this result, we can also prove quite quickly that


(ABC)T = CT BT AT .

(148)

If I take my equation (146) above and transpose it, then I get


(~ATi K~A j )T = 2j (~ATi M~A j )T ,

(149)

and using the above theorem of matrix algebra gives


~ATj K~Ai = 2j ~ATj M~Ai ,
39

(150)

remembering that the transpose operation twice just goes back to the original matrix and that since
M and K are symmetric the transpose operation has no effect on them. Subtracting equations (145)
and (150) gives me the interesting result

It follows that

(2j 2i )~ATj M~Ai = 0 .

(151)

~ATj M~Ai = 0 and ~ATj K~Ai = 0 ifi 6= j .

(152)

These equations describe a property of the vectors ~A j called orthogonality with respect to the matrices M and K. This property turns out to be very useful.
For the case i = j, we define by
and

~ATi M~Ai = Mii

(153)

~ATi K~Ai = Kii

(154)

the generalised mass and stiffness. We will now make use of these ideas in a very powerful method.
Before we can do so, we must make one last definition; an object called the modal matrix P. We
define it to be the sequence of all the column vectors ~Ai , i.e.
P = [~A1~A2 . . . ~An ] ,

(155)

so that P is a n n matrix formed by the eigenvectors ~Ai . What is the use of this matrix? Consider
the matrix
(156)
PT MP = [~A1~A2 . . . ~An ]T M[~A1~A2 . . . ~An ] .
Because of the orthogonality relations above, most of the terms when we multiply out this expression are zero. We are left with

M11
0 ... 0
0 M22 . . . 0

PT MP =
(157)
.
.
.
0
0 . . . Mnn
which is a diagonal matrix. The matrix PT KP is also diagonal. The result we have proved is that
the modal matrix diagonalises the mass and stiffness matrices. This is useful! If we consider the
equation of motion
M~x + K~x = ~F0 exp(it) ,
(158)
then if we define a new set of variables by ~x = P~y then
MP~y + KP~y = ~F0 cos(t) ,

(159)

and multiplying this equation by PT gives


PT MP~y + PT KP~y = PT ~F0 cos(t) .
40

(160)

Since the matrices are diagonal, this uncouples the problem: This equation is just n single degree of
freedom problems
Mii yi + Kii yi = [PT ~F0 ]i cos(t)
(161)
which we know how to solve.
The above mathematics seems quite complicated, but dont be intimidated. You can try the method
for yourself. In brief, to solve any forced multiple degree of freedom problem you need to:
Draw a free body diagram
Write down the equation of motion for each component
Collect the equations together in matrix form
Set the forcing to zero temporarily and solve the freely vibrating system to find the eigenvalues
(i.e. frequencies of free vibration)
Use the eigenvalues to solve the matrix equation and determine the eigenvectors
Construct the modal matrix P from the eigenvectors
Use the modal matrix to define a new set of coordinates yi (this is conceptual - you dont need
to calculate anything here)
Calculate the products PT MP and PT KP and write down the new matrix equation for y
Expand out the matrix equation to give n single degree of freedom systems
Solve the single degree of freedom systems
Use ~x = P~y to put the solution back into the original coordinates
Its quite a lengthy procedure, but the steps are quite simple to carry out if you have a computer at
hand - and it does give a general method to solve any system, however complicated, provided you
can calculate the eigenvalues and vectors. In particular, the method is suited to programs such as
MATLAB where you can define and handle matrices easily.

41

4 Applications
4.1 Accelerometer
As a first application of the theory of vibration, we consider a device known as an accelerometer.
As the name suggests, it is used to measure the acceleration of a vibrating surface. Essentially,
the device consists simply of a seismic mass, basically a lump of metal, that is attached by a
spring damper system to the surface whose vibration is to be measured. A sensor is able to measure
the distance between the seismic mass and the surface. Our task is to relate the vibration of the
mass, which we can measure, to the acceleration of the surface. Figure 21 shows a diagram of the
situation.

c
m
k
y (t)

Figure 21: A schematic diagram of an accelerometer. x measures distance from a fixed point in
space to the mass, y measures distance from a fixed point in space to the vibrating surface.
The equation of motion is
mx + c(x y)
+ k(x y) = 0 .

(162)

We now introduce the new variable z = x y. We can differentiate to find


z = x y ,

(163)

z = x y ,

(164)

and then eliminate x from the equation of motion in favour of z. We get


mz + cz + kz = my ,

(165)

which looks quite familiar. Let us assume that the surface is vibrating with frequency so that
y(t) = Y cos(t)
42

(166)

and so
y = 2Y cos(t) ,

(167)

which we can use in our equation of motion to give


mz + cz + kz = m2Y cos(t) ,

(168)

which is in the form which we are used to solving, with the amplitude of the forcing given by m2Y .
Dividing through by m we have
z + 20 z + 20 z = 2Y cos(t) ,

(169)

We know how to solve this sort of equation. The canonical form is


z + 20 z + 20 z = 2 z0 cos(t) ,

(170)

By comparison, we find that the static response z0 is


z0 = Y 2 /20

(171)

and the amplitude of the dynamic response is


z0
Y (/0 )2
Z=p
=p
.
(1 (/0 )2 )2 + 42 (/0 )2
(1 (/0 )2 )2 + 42 (/0 )2

(172)

Hence, the ratio of the amplitudes of vibration of the accelerometer and the surface is
(/0 )2
Z
=p
.
Y
(1 (/0 )2 )2 + 42 (/0 )2

(173)

If we make the free vibration frequency 0 very high by a suitable choice of k and m, then /0
will be small and
Z
2
= 2
(174)
Y
0
to a good approximation. Therefore we can write
2Y = 20 Z .

(175)

Now, the surface is moving sinusoidally according to


y = Y cos(t)

(176)

y = 2Y cos(t)

(177)

A = 2Y

(178)

and so the acceleration of the surface is

and the amplitude of this acceleration is

43

Therefore,
A = 20 Z .

(179)

Since we can measure the amplitude Z and we already know the frequency 0 , we can measure the
acceleration amplitude A. Since we are far below the free vibration frequency, our measured signal
z(t) will be in phase with the term on the right-hand side of Eq. (169) and therefore in antiphase
with the acceleration.
A further degree of sophistication is to add damping to the accelerometer in order to improve its
accuracy. The whole idea in going from Eq. (173) to (174) is that the term under the square root
should be as close as possible to 1. This is achieved by making /0 as small as possible. To do
even better, we retain an extra term of order 2 /20 :
Z
Y

(/0 )2
p
(1 (/0 )2 )2 + 42 (/0 )2

2
[1 (22 1)(/20 )] .
20

(180)

This time we have kept extra terms to which is why our original result (174) is not exact. These
2
2
terms become
significant when /0 is not so small. However, we can get rid of them! If we
set = 1/ 2 then the above equation reduces to equation (174). We can therefore expect our
accelerometer to be more accurate if we apply damping so that 0.7.
Many accelerometers work on this principle. Low frequency accelerometers use a damping ratio
of 0.7 as described above. This also improves the phase distortion; you can read a description of
phase distortion in Thompsons book if youre interested. The ones you will meet in the lab use a
piezoelectric crystal which has a very high natural frequency and therefore there is no need for any
damping (can you explain why?). These accelerometers are more suited to high frequency work.

4.2 Shaft whirling


The next system we will look at is the vibrations of a rotating out-of-balance shaft. For simplicity,
lets consider a massless shaft with a disc of mass m in the centre. Figure 22 shows what this looks
like.
Figure 23 shows the situation we will analyse. The origin of the coordinates has been chosen to
be at the equilibrium position of the shaft centre. The shaft centre is attached to the disc at C, so
in equilibrium C rests at O. The centre of mass G is rotated about the shaft centre C at the same
angular velocity as the rotation of the shaft, which is .

44

w
C

Figure 22: Schematic diagram of a disc rotating on a shaft. The shaft AB meets the disc at C. The
centre of mass of the disc is located at G.
Because the shaft is rotating, C can be displaced from its equilibrium position and will then be
rotated about O. We want to know how far the centre of the shaft C is displaced from its equilibrium
position.

G
e

w t

z
C

q
x

Figure 23: Motion of the shaft centre C and the centre of mass G. G moves around C due to the
rotation of the shaft. C moves around O because of the inertial forces caused by rotating G.

45

Our first step in analyzing this problem is to draw the free body diagram. This is shown in Figure
24.

m e w

m x

G
:

m y
k z
C

x
O

Figure 24: Free body diagram for the system of Figure 22


The force acting on C is due to the displacement of the centre of the shaft from its equilibrium
position. If the force for a unit displacement is k then the force here is kz directed towards O. The
value of k is a property of the shaft and the way it is mounted and can be calculated using solid
mechanics. The force acting on G is the centrifugal force due to its rotation about C. This is the
forcing in our problem, which causes C to be displaced from O. The other two forces shown acting
on G are the inertial forces due to its motion. Note that we must give two of these since there are
two independent coordinates needed to specify the position of G. G and C are rigidly linked so the
equation of motion can be deduced by summing all the forces in the system. We get two equations,
by resolving in the x and y directions:
mx + kz cos() me2 cos(t) = 0 ,
2

my + kz sin() me sin(t) = 0 .

(181)
(182)

Now x = z cos() and y = z sin(), so we can write the two equations of motion
mx + kx = me2 cos(t) ,
2

my + ky = me sin(t) .

(183)
(184)

This is a most simple two-degree of freedom system since the two equations are not coupled. Therefore each of them can be solved using the usual method for forced SDF systems. We find
x(t) = e

2 /20
cos(t) ,
1 2 /20
46

(185)

y(t) = e

2 /20
sin(t) .
1 2 /20

(186)

These can be added together to calculate z2 = x2 + y2 . We find


z=e

2 /20
.
1 2 /20

(187)

where 0 is the natural frequency of free vibration for the shaft. If is small, i.e. the shaft rotates slowly, then k m2 and z = e(/0 )2 , which shows C is not far displaced from O. At
higher speeds of rotation, however, C does move a long way from O and at resonance ( = 0 ) the
expression goes to infinity. At high speeds, for which is large,
z e ,

(188)

so the centre of mass sits at the origin and the shaft centre C rotates about it.
This simple example shows that shaft whirling occurs at the natural frequency of vibration of the
shaft. This is a very useful result because now we can calculate when whirling will occur by analyzing the free vibrational properties of the shaft.

4.3 Beating
We will look at the beating behaviour of two pendulums. This will allow us a bit of practice on
inserting the initial conditions into a MDF system as well as providing an interesting study of the
phenomena of beating.

a
q

k
q

Figure 25: Two weakly coupled pendulums.

47

Beating occurs when we have a weak link between two systems with natural frequencies that are
close together. Im going to consider two identical pendulums joined by a weak spring. In reality
the coupling between the pendulums might occur because of the motion of the mounting plate, but
we will just take it as given here, and model it using the weak spring. Figure 25 shows the two
pendulums and Figure 26 is the free body diagram.

k a (q 2- q 1)
q
1

m l2q

m l2q

m g

m g

Figure 26: Free body diagram for the system of Figure 25.
From the free body diagram, we can write down the equation of motion by taking the moments of
the forces about the pivots of the pendulums.
ml 2 1 + mgl sin 1 ka2 (2 1 ) = 0 ,
ml 2 2 + mgl sin 2 + ka2 (2 1 ) = 0 .

(189)
(190)

In the approximation of small oscillations, we can put sin and write the equations of motion
in matrix form:

ml 2 0
1
mgl + ka2 ka2
1
+
=0 .
(191)
0
ml 2
2
ka2
mgl + ka2
2
We substitute ~ = ~A exp(it). If we carry out the differentiation on t and insert the result in the
equation of motion we get

ml 2 2 + mgl + ka2 ka2


A1
=0 .
(192)
ka2
ml 2 2 + mgl + ka2
A2
For the equation to have solutions, the determinant of the matrix must be zero which gives us our
usual quadratic equation for 2 . If we do the algebra we get

48

ml 2 2 + mgl + ka2 = ka2 .

(193)

It follows that the values of are


r
+ =

r
g
l

g 2ka2
.
+
l
ml 2

(194)

Putting these values back into equation (192) gives the two equations for the mode shapes which
are

1
1
A1
1 1
A1
2
2
ka
= 0 , ka
=0 .
(195)
1 1
A2
1 1
A2
The mode shapes for A are therefore

~A+ =

1
1

~A =

1
1

(196)

Now we will construct the general solution.


The two solutions we have determined, which are

~ = c+ 1 exp(i+t)
1
and

~ = c

1
1

(197)

exp(it)

(198)

are just the complementary functions. I have put in the factors of Q+ and Q because the mode
shapes are only determined up to an arbitrary constant. Remember, this constant can be complex!
Dont be put off by the fact that there are vectors - they behave in just the same way as the complementary functions of the SDF systems we looked at before. The general solution is just the sum of
the complementary functions. It is

~(t) = Re c+ 1 exp(i+t) + c 1
exp(it) .
(199)
1
1
At this point in the calculation, I will take the real part, writing Re[c+ exp(i+t)] = B sin(+t) +
C cos(+t) and Re[c exp(it)] = D sin(t) + E cos(t). This gives

1
B sin(+t) +C cos(+t) + D sin(t) + E cos(t)
=
.
(200)
2
B sin(+t) +C cos(+t) D sin(+t) E cos(+t)
This then is the general solution, and it contains four unknowns B,C, D, E that are determined by the
initial conditions. For the initial condition, I will specify that both pendulums are at rest, pendulum
one is at an angle to the vertical and pendulum two is vertical. If we put these conditions into the
49

above equations for 1 and 2 , we find C = E = 0 and B = D = /2. Please check this for yourself.
The final solution is therefore
1 = (/2)[cos(1t) cos(2t)]

(201)

2 = (/2)[cos(1t) + cos(2t)]

(202)

and
We can make use of the trig identities to write this in a more convenient form. We have
cos A + cos B = 2 cos

A+B
AB
cos
2
2

cos A cos B = 2 sin

A+B
AB
sin
2
2

so our solution can be written as


1 = cos((1 + 2 )t/2) cos((1 2 )t/2)

(203)

2 = sin((+ + )t/2) sin((+ )t/2)

(204)

and
Finally we can put in the values we calculated for + and . If the coupling is very weak, k is
small and so
r
+ +
g

(205)
2
l
and
+ 1
=
2
2

If we write

g
1
l

r
=

g ka2
l 2mgl

r
g ka2

1+

mgl
l 2mgl
2ka2

r
,

g
,
l

(206)

(207)

then the solution can be written as


1 = cos(t) cos(t) ,

(208)

2 = sin(t) sin(t) .

(209)

A plot of this form is shown in Figure 27. As you can see, the vibration amplitude is periodically
modulated, and the energy of vibration oscillates between the two pendulums with a period of /.
This phenomenon is called beating.

50

1.0

1/

0.5
0.0

-0.5
-1.0
1.0

2/

0.5
0.0

-0.5
-1.0

2/
2/

Figure 27: Plot of the beating behaviour in the coupled pendulum system. The outer line shows the
sinusoidal envelope curve

51

F (t)

x
1

F (t)
m
1

x
k

x
2

k
1

Figure 28: Left: a forced oscillator. Right: the same oscillator with an additional spring-mass
system acting as antivibration device.

4.4 Anti-resonance and vibration absorbers


We have seen in previous sections that the vibration of a forced system can be always suppressed if
one makes the natural frequency low enough, i.e. by designing a system with large mass and small
stiffness. However, this may often not be a feasible option due to other design constraints.
In this case, another strategy may be chosen to suppress unwanted oscillations. Consider the system
in Figure 28 (left) with a sinusoidal force F(t) = F0 cos(t). The equation of motion of this system
is
m1 x1 + k1 x1 = F0 cos(t)
(210)
and the steady-state response of the system is
x(t) =

x0
F0
cos(t) =
cos(t) .
k1 m1 2
1 2 /20

If the forcing frequency is not much larger than the systems natural frequency 0 =
response amplitude may be significant, and at resonance it becomes infinite.
52

(211)
p

k1 /m1 the

How can we suppress the vibration of the mass m1 without changing the mass or the stiffness of the
system? We will see that we can deal with this problem by adding another mass-spring subsystem
(m2 and k2 ) as shown in Figure 28 (right). This gives a two-degree of freedom system
m1 x1 + k1 x1 + k2 (x1 x2 ) = F0 cos(t)

(212)

m2 x2 + k2 (x2 x1 ) = 0

(213)

or, in matrix form

x1
F0
k1 + k2 k2
m1 0
x1
=
cos(t) .
+
0
x2
k2
k2
x2
0
m2
We solve this by assuming

x1
x2

X1
X2

(214)

cos(t) .

(215)

where X1 and X2 are the vibration amplitudes of the masses m1 and m2 . This leads to the system of
equations
(m1 2 + k1 + k2 )X1 k2 X2 = F0 ,
k2 X1 + (m2 + k2 )X2 = 0 .
2

(216)
(217)

By solving this system we obtain the vibration amplitudes X1 and X2 . The result is
X1 =
X2 =

(k2 m2 2 )F0
,
K()
k2 F0
,
K()

(218)
(219)

where
K() = (k1 + k2 m1 2 )(k2 m2 2 ) k22 .

(220)

We see that the amplitude of vibration of the mass m1 becomes zero if k2 and m2 fulfil the relation
k2 /m2 = 2 , i.e., if the resonant frequency of the added subsystem matches the excitation frequency.
p
In particular, we can suppress the resonant response of the original system at = 0 = k1 /m1
completely by chosing k2 /m2 = k1 /m1 .

4.5 Out of balance in reciprocating internal combustion engines


The inertial forces due to the motion of the piston and con rod in an engine must be balanced by
forces acting on the engine and these can cause vibration. In this section we will deduce what these
forces are and what measures can be taken to keep them to a minimum. Figure 29 shows a schematic
diagram of a piston in a cylinder a distance x from bottom dead centre attached by a con rod AB to

53

f
A

l
l+ r -x

r
C

Figure 29: Schematic diagram of a piston, con rod AB and crank BC.
a crank shaft BC which rotates about C. The length of the shaft is r and the length of the con rod is
l. The distance between the piston and the crank axis is therefore r + l x as shown.
To find the inertial force due to the motion of the piston, con rod and crank shaft, we will seek to
relate x to , the angle through which the crank shaft has rotated. If the engine runs at a constant
the angular velocity of the crankshaft, is a constant .
speed then ,
The problem is that because of the geometry the motion x(t) of the piston is not simply a sinusoidal
function. To work out what it is, we project the lengths of AB and BC onto AC to give
AB cos + BC cos = AC

(221)

and putting in the various terms gives us


l cos + r cos = l + r x

(222)

sin sin
=
l
r

(223)

From the sine law, we have


therefore we can eliminate because

q
cos = 1 sin2 =
hence

s
1

r2 sin2
.
l2

(224)

s
x = r + l r cos l

r2 sin2
.
l2

If we introduce a parameter = r/l then we have

1p
1
2
2
1 sin
.
x = r 1 + cos

54

(225)

(226)

In an engine, the value of is usually around 1/3 to 1/4 and so we can expand the square root as a
power series and ignore higher powers of because they will be small. For small , we have from
the Taylor series that
p
2 sin2
1 2 sin2 1
,
(227)
2
so we can write

sin2
x = r 1 cos +
.
(228)
2
where I have ignored powers of higher than one. This approximation is usually good enough. We
now want to differentiate twice with respect to t in order to find the acceleration. We use the chain
rule so
d
sin = cos = cos ,
(229)
dt
where is the constant angular velocity described above. Carrying out this procedure on x gives
x = r(sin + sin cos ) = r(sin +

sin 2) ,
2

(230)

and
x = r(cos + cos 2)2 .

(231)

Now we have calculated the acceleration of the piston, let us consider the forces acting in the system
of Figure 29. To make the analysis easier, we will model the con rod in an approximate way. The
true con rod shape is rather complicated, and we will replace it in our lumped model with two
masses, m1 and m2 , one at each end of the real rod. Figure 31 shows the con rod and its model.

Figure 30: Schematic diagram of the con rod and its model. G is the centre of gravity.

55

We choose the lengths a and b and the masses m1 and m2 so that m1 + m2 = MCR , where MCR is the
mass of the con rod, and m1 a = m2 b so the the centre of gravity of the con rod and the model of the
con rod are the same. Figure 30 shows the forces acting on the system with the model for the con
rod inserted. In order to balance the system, the size and shape of the crank is chosen so the rotation
of the mass labelled m3 causes the same centrifugal force as the rotation of the mass m2 from the
con rod. These forces balance, but the remaining horizontal component of the inertial force M1 x,

where M1 = MP + m1 and MP is the mass of the piston, causes a force H on the crankshaft and hence
on the engine.

M
p

rW

m
1

m
2

rW
2

Figure 31: Forces acting on the piston, con rod and crank.
To summarise this section, we have looked at the motion of the piston, con rod and crankshaft
system at a constant angular velocity and deduced the inertial force due to the motion of the piston
and the con rod. The crankshaft shape is chosen to eliminate the non-horizontal component of the
con rod inertial force and the resulting force on the engine is given by
H = (MP + m1 )x = (M1 )r(cos + cos 2)2

(232)

The term M1 r2 cos is called the primary unbalance force and the term M1 r2 cos 2 is the secondary unbalance force. To understand the origin of the secondary unbalance force, it is useful
to have a look at Figure 32. The motion x() of the piston is not strictly sinusoidal, and the difference with a pure cosine can be well approximated by a cosine with twice the frequency. The
corresponding acceleration gives rise to the secondary unbalance force.
We now look at the assembly of cylinder in an engine, and the resultant forces which act on the body
of the engine. The goal will be to devise an engine that is balanced, so the unbalance forces from
each cylinder cancel out. We will consider a four cylinder inline arrangement. Other arrangements
can be analysed by the same method.
Figure 32 illustrates a four cylinder engine where the pistons are connected to the crankshaft at
angles of 0, , and 0 radians. We want to know what the forces and moments acting on the engine
56

2.0

x/r

1.5

1.0

0.5

0.0
0

10

Figure 32: Upper full line: motion of the piston as a function of the angle of rotation of the
crankshaft. Dotted line: approximation by a cosine. Lower full line: Difference between the actual
motion and a cosine.
are. We will consider a general case, then see how this applies to the four cylinder engine.
Suppose that there are N cylinders in a line, and that the i-th crank is offset by an angle of i with
reference to the vertical (it doesnt have to be the vertical but it may as well be. As long as we
use the same reference point for each cylinder everything will work out ok). After time t, the shaft
rotates through an angle t and the crank is at an angle i t to the vertical. The force due to the
piston and con rod is
Hi = M1 r2 (cos(i t) + cos 2(i t))
(233)
The total force is the sum over all the cylinders, so
H = Hi

(234)

and is composed of a primary and secondary part. I will call FP the primary and FS the secondary
unbalance forces. They are given by
FP = M1 r2 cos(i t) = M1 r2 [cos t cos i + sin t sin i ]
i

57

(235)

Figure 33: Four cylinder inline arrangement. The cranks are attached at angles of 0, 180, 180 and 0
degrees.
and
FS = M1 r2 cos 2(i t) = M1 r2 [cos 2t cos 2i + sin 2t sin 2i ] .
i

(236)

We also want to know the moments acting on the engine. If we take moments about a point O, and
each crank i is a distance zi from O, the the primary and secondary moments are
FP = M1 r2 zi cos(i t) = M1 r2 [cos t zi cos i + sin t zi sin i ]
i

(237)

and
FS = M1 r2 zi cos 2(i t) = M1 r2 [cos 2t zi cos 2i + sin 2t zi sin 2i ] . (238)
i

We can now evaluate the primary and secondary moments and forces from these expressions for
any inline arrangement of cylinders. For the four cylinder example, we have i = 1, 2, 3, 4 and z1 =
z4 = a, z2 = z3 = b. We can evaluate the sums of the various terms as shown in Table 1. The terms
in the columns are the terms in the sums in the equations above.
Using the table we can see that all the moments, and the primary force unbalance cancel, but a
secondary unbalance force remains. This force can be countered using an out-of-balance rotating
shaft which is geared to rotate at twice the speed of the engine. This adds to the complexity of the
engine, but ensures that it does not vibrate excessively.

58

crank
1
2
3
4
total

i
0

zi
a
b
-b
-a

sin i
0
0
0
0
0

crank
1
2
3
4
total

i
0

zi
a
b
-b
-a

zi sin i
0
0
0
0
0

cos i
1
-1
-1
1
0

sin 2i
0
0
0
0
0

zi cos i
a
-b
b
-a
0

cos 2i
1
1
1
1
4

zi sin 2i
0
0
0
0
0

zi cos 2i
a
b
-b
-a
0

Table 1: Sums for the primary and secondary unbalance forces and moments in a four cylinder
inline engine as shown in Figure 28
Exercise: 27
Another solution to the problem of minimising engine vibration is to use 6 inline cylinders at angles of 0o , 120o , 240o , 240o , 120o and 0o and at distances of
2.5, 1.5, 0.5, 0.5, 1.5, 2.5 from the centre of the shaft. Draw your own version of Table 1 and confirm that the primary and secondary unbalance forces and
moments all cancel. The six cylinder arrangement is therefore very common.

59

Potrebbero piacerti anche